Organophosphate Ester, 2-Ethylhexyl Diphenyl Phosphate (EHDPP

DOI: 10.1021/acs.est.8b06246. Publication Date (Web): January 17, 2019. Copyright © 2019 American Chemical Society. Cite this:Environ. Sci. Technol. ...
0 downloads 0 Views 872KB Size
Subscriber access provided by Iowa State University | Library

Ecotoxicology and Human Environmental Health

Organophosphate Ester, 2-Ethylhexyl Diphenyl Phosphate (EHDPP), Elicits Cytotoxic and Transcriptomic Effects in Chicken Embryonic Hepatocytes and its Biotransformation Profile Compared to Humans Jinyou Shen, Yayun Zhang, Nanyang Yu, Doug Crump, Jianhua Li, Huijun Su, Robert J. Letcher, and Guanyong Su Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b06246 • Publication Date (Web): 17 Jan 2019 Downloaded from http://pubs.acs.org on January 18, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34

1

Environmental Science & Technology

Table of Content (TOC) Graphical Art

2 3 4

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 34

5

Organophosphate Ester, 2-Ethylhexyl Diphenyl Phosphate (EHDPP), Elicits Cytotoxic and

6

Transcriptomic Effects in Chicken Embryonic Hepatocytes and its Biotransformation Profile

7

Compared to Humans

8

Jinyou Shen†,#, Yayun Zhang†,#, Nanyang Yu‡, Doug Crump§, Jianhua Li†, Huijun Su†, Robert J. Letcher§,

9

Guanyong Su†,*

10 11

† Jiangsu

12

Biological Engineering, Nanjing University of Science and Technology, Nanjing 210094, P. R. China

13



14

University, Nanjing 210023, China

15

§

16

Research Centre, Carleton University, Ottawa, ON, K1A 0H3, Canada

Key Laboratory of Chemical Pollution Control and Resources Reuse, School of Environmental and

State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing

Ecotoxicology and Wildlife Health Division, Environment and Climate Change Canada, National Wildlife

17 18

* Corresponding authors:

19

Tel.: 86-1-395-176-3661,

20

E-mail: [email protected],

21

#

These two authors contribute this work equally.

22

ACS Paragon Plus Environment

Page 3 of 34

23 24

Environmental Science & Technology

Abstract The effects of 2-ethylhexyl diphenyl phosphate (EHDPP) on cytotoxicity and mRNA expression, as

25

well as its metabolism, were investigated using a chicken embryonic hepatocyte (CEH) assay. After

26

incubation for 36 h, the lethal concentration 50 (LC50) was 50 ± 11 μM, suggesting that EHDPP is one of a

27

small cohort of highly toxic organophosphate esters (OPEs). By use of a ToxChip polymerase chain reaction

28

(PCR) array, we report modulation of 6, 11 or 16/43 genes in CEH following exposure to 0.1, 1 or 10 μM

29

EHDPP, respectively. The altered genes were from all 9 biological pathways represented on the ToxChip

30

including bile acids/cholesterol regulation, glucose metabolism, lipid homeostasis and the thyroid hormone

31

pathway. After incubation for 36 h, 92.5 % of EHDPP was transformed, and one of its presumed

32

metabolites, diphenyl phosphate (DPHP), only accounted for 12% of the original EHDPP concentration.

33

Further screening by use of high-resolution spectrometry revealed a novel EHDPP metabolite, hydroxylated

34

2-ethylhexyl monophenyl phosphate (OH-EHMPP), which was also detected in a human blood pool.

35

Additional EHDPP metabolites detected in the human blood pool included EHMPP and DPHP. Overall, this

36

study provided novel information regarding the toxicity of EHDPP and identified a potential EHDPP

37

metabolite, OH-EHMPP, in both avian species and humans.

38

Keywords: organophosphate ester, 2-ethylhexyl diphenyl phosphate (EHDPP), lipid homeostasis, avian

39

species, humans

40

ACS Paragon Plus Environment

Environmental Science & Technology

41

Page 4 of 34

Introduction

42

The production and use of brominated flame retardants (BFRs) such as polybrominated diphenyl ethers

43

(PBDEs) and hexabromocyclododecane (HBCDD) has been restricted under the Stockholm Convention. As

44

a result, the market share of organophosphate ester (OPE) flame retardants (FRs) has been increasing

45

significantly in recent years.1 In Europe, the total consumption of OPE-FRs in 2006 was approximately 90,000

46

tonnes (equivalent to 20% of the total consumption) and the current market share of OPE-FRs is even greater.2

47

In China, the annual consumption of OPE-FRs was approximately 70,000 tons in 2007 with an annual

48

estimated increase of 15 %.3 Thus, there is great concern regarding the occurrence, transformation and health

49

effects of OPE-FRs in the environment.4, 5

50

2-Ethylhexyl diphenyl phosphate (EHDPP) is one typical OPE with a phosphate ester backbone

51

structure.2 EHDPP is used widely in various products including, polyvinyl chloride (PVC), rubber, and food

52

packaging, as a plasticizer and a FR.6,

53

products and therefore can easily leach out into the environment during manufacture, usage, and disposal of

54

products.8

7

As an additive to polymers, EHDPP is not chemically bound to

55

EHDPP is ubiquitous in various environmental samples, such as indoor air, dust, surface water, wildlife

56

and even humans.8-12 For instance, in indoor dust from Portugal, Denmark, the U.K., and the U.S., EHDPP

57

was a dominant contaminant, with concentrations up to 1.1 μg/g, 0.79 μg/g, 1.6 μg/g, and 3.3 μg/g,

58

respectively.13-15 Kim et al investigated the OPEs in a wastewater treatment plant in New York, and found

59

that EHDPP was one of the dominant compounds in ash samples with a concentration of 288 ng/g dry weight.10

60

Two recent studies have reported detection frequencies of EHDPP in surface water samples from Taihu Lake

61

in China were 88 % and 100 % with mean concentrations of 1.9 ng/L and 2.8 ng/L, respectively, which could

62

pose a risk to local aquatic organisms.16, 17 Notably, EHDPP concentrations in biota were higher than in surface

63

water from the same sampling sites. In a Swedish lake and Manila Bay (Philippines), concentrations of 14 and

64

2.1 μg EHDPP/g lipid weight in fishes were reported, respectively.18, ACS Paragon Plus Environment

19

In a recent study, Zhao et al

Page 5 of 34

Environmental Science & Technology

65

investigated human prenatal exposure to OPEs by analyzing chorionic villus samples; EHDPP was detectable

66

in 100% of analyzed samples with a median concentration of 13.6 ng/g dry weight.9 These extremely high

67

detection frequencies warrant a comprehensive assessment of potential adverse effects to exposed organisms.

68

Data regarding the toxicity of EHDPP are scarce, but a few studies have shown that EHDPP can induce

69

adverse effects such as endocrine disruption and developmental and neurotoxic effects in exposed

70

organisms.20-22 Specifically, recent studies have investigated the activation of human peroxisome proliferator-

71

activated receptor gamma (PPARγ) and estrogen-related receptor γ (ERRγ) using luciferase reporter gene

72

assays and demonstrated that EHDPP bound to and activated PPARγ and ERRγ.20, 21 Schang et al investigated

73

the possible activity of EHDPP as an endocrine disruptor in MA-10 mouse Leydig tumor cells, and reported

74

that EHDPP altered the expression of genes involved in mitochondrial activity, cell survival, superoxide

75

production, and basal or stimulated steroid secretion.22 Behl et al evaluated a set of OPEs, including EHDPP,

76

using multiple cell-based in vitro assays and demonstrated that EHDPP (1-10 μM) exhibited comparable

77

developmental and neurotoxic effects as two BFRs, 3,3′,5,5′-tetrabromobisphenol A (TBBPA) and 2,2′4,4′-

78

brominated diphenyl ether (BDE-47).23 Overall, EHDPP toxicity data are still unavailable for most organisms,

79

including avian species, and more comprehensive studies are warranted to understand its underlying

80

mechanism(s) of toxicity.

81

In terms of biological transformation of EHDPP, an in vitro study using human liver microsomes

82

suggested two crucial pathways: 1) hydroxylation of the phenyl ring; and 2) O-dealkylation between the 2-

83

ethylhexyl functional groups and phosphate acid. These pathways would result in the formation of diphenyl

84

phosphate (DPHP) and hydroxylated EHDPP (OH-EHDPP).24 DPHP has been frequently reported in human

85

urine samples and its concentrations were positively correlated with thyroid hormones (T4).25-27 However,

86

DPHP is not a specific metabolite for EHDPP because it can be derived from multiple parent compounds such

87

as triphenyl phosphate (TPHP), resorcinol bis(diphenylphosphate) (RDP), and isopropylated and tertbutylated

88

triarylphosphate esters (ITPs & TBPPs).28-31 OH-EHDPP appears to be an unique biomarker of EHDPP, but ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 34

89

concentrations of OH-EHDPP in human samples are very low (i.e. 0.09 ng/mL urine).32 Investigating the

90

metabolism of EHDPP and identifying potential biomarkers is critical for understanding its adverse effects to

91

organisms. Currently, information on the metabolism of EHDPP remains unknown for avian species, and

92

differences of biotransformation between avian species and humans are also unknown.

93

The present study investigated the cytotoxicity and mRNA expression levels of 43 genes from 9 pathways

94

– bile acids/cholesterol regulation, cell cycle, DNA repair, glucose metabolism, immune response, oxidative

95

stress, the thyroid hormone pathway, lipid homeostasis, and xenobiotic metabolism – in chicken (Gallus gallus

96

domesticus) embryonic hepatocytes (CEHs) following exposure to EHDPP. Additionally, the concentration

97

of EHDPP and DPHP in cultured cells and medium were quantified to characterize the mass balance between

98

EHDPP and DPHP and cell accumulation. Finally, other potential metabolites besides DPHP were identified

99

by use of high-resolution quadrupole-time of flight mass spectrometry (Q-ToF-MS), and newly identified

100

metabolites were further screened in a pool of human blood from the Chinese population to examine the

101

difference of EHDPP metabolism between an avian in vitro assay (CEHs) and humans.

102 103

Experimental section

104

Chemicals

105

All chemicals, i.e. EHDPP (CAS No.: 1241-94-7), DPHP (CAS No.: 838-85-7), and tris(1,3-dichloro-2-

106

propyl) phosphate (TDCIPP; CAS No.: 13674-87-8), were purchased from Sigma-Aldrich (St. Louis, MO,

107

U.S.A.). The purity of these chemicals was reported to be greater than 99% by manufacturer.

108 109

Preparation of Chicken Embryonic Hepatocytes (CEH)

110

Detailed information for the preparation of CEHs can be found elsewhere.33, 34 In brief, we purchased

111

fertilized and unincubated white leghorn chicken eggs from the Canadian Food Inspection Agency , and these

112

eggs were incubated (37.5 °C, 60 % relative humidity) for 19 days. After incubation for 19 days, embryos ACS Paragon Plus Environment

Page 7 of 34

Environmental Science & Technology

113

were decapitated and livers were removed and pooled rapidly. For cell separation and reduction of cell

114

aggregation respectively, hepatocytes were treated with Percoll and DNase I in sequence. After centrifugation,

115

cells were then suspended by use of 32 mL of Medium 199 that were supplemented with 1μg/mL of insulin

116

and thyroxine. Once the cells were distributed into each well of 48-well plate, the hepatocytes were incubated

117

for 24 h under conditions of 37.5 °C and 5% CO2. After 24 h, the CEHs were dosed to DMSO (2.5 μL/well;

118

n = 3) and EHDPP (nominal concentrations: 100, 20, 10, 2, 1, 0.1 μM for cell viability; 10, 1, and 0.1 μM for

119

variation of mRNA levels; 1 μM for in vitro metabolism; n = 3 for all treatments) and incubated for additional

120

hours (i.e. 0, 12 or 36 h) depending specific purposes.

121 122

Cell viability

123

ViaLightTM Plus BioAssay kit was used for the measurement of adenosine triphosphate (ATP) in CEHs

124

exposed to EHDPP for 36h. Based on our previous studies, TDCIPP at a nominal concentration of 300 μM

125

was used as a positive control.30, 35 Following the 36h incubation period, the plates were maintained at 20 oC

126

for no less than 5 min. Then, CEH medium was gently aspirated from each well of the plates, and fresh medium

127

and cell lysis reagent was successively added into each well. After mixed well, a 100 μL aliquot of this

128

medium/lysis reagent mixture was transferred to a luminometer plate, and incubated for 2 min at 20 oC. The

129

mixture was then mixed with an aliquot of 100 μL of ATP monitoring reagent plus, and immediately measured

130

for intensity of luciferase luminescence by use of the Luminoskan Ascent.

131 132

Variation of mRNA levels in CEH following exposure to EHDPP

133

EHDPP was investigated at three concentrations (10, 1 and 0.1 μM) to determine effects on gene

134

expression. After exposure for 36 h, we further examined variation of mRNA levels in CEH by use of the 4th

135

generation custom chicken RT2 Profiler PCR Array (ToxChip) built in a 96-well plate.34 The 96-well ToxChip

136

contained 43 target genes and 5 control genes in duplicate. Biological functions of these 43 target genes were ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 34

137

described in Table S1, and the 5 control genes involved 1 positive PCR control (PPC), 1 reverse transcription

138

control (RTC), 2 housekeeping genes, and 1 genomic DNA control, respectively.

139

Isolation of total RNA from CEH were detailed in our previous publications.30, 35 In brief, plates were

140

removed from the -80 oC freezer, and the cell lysis reagent was added into each well of plates. Then, 200 ng

141

of RNA was reverse-transcribed to complimentary DNA (cDNA), that was gently mixed with the RT2 SYBR

142

Green Mastermix (Qiagen). A 25 μL aliquot of this mixture was immediately added to each well on the 96-

143

well ToxChip. All arrays were measured in the Agilent Stratagene MX3005 (Agilent Technologies, Santa

144

Clara, CA, U.S.A.) using the thermal profiles reported in our previous publications.30,

145

measurement of PCR arrays, amplification of gene expression was not observed for the genomic DNA

146

contamination control (i.e. Ct value ≥ 35), and PPC and RTC controls met the appropriate QA/QC criteria (i.e.

147

ΔCt ≤ 5; ΔCt = Ct values of RTC – Ct values of PPC). These results ensure robust gene expression analysis.

148

Importantly, the relative cycle thresholds of the two housekeeping genes were invariable (ANOVA, p > 0.05)

149

regardless of treatment with EHDPP at any of the administered concentrations.

35

We During the

150 151

In Vitro Metabolism of EHDPP

152

A non-cytotoxic concentration of EHDPP (1 μM; see Cell Viability section) was administered to CEH

153

for the in vitro metabolism evaluation. DMSO-treated cells were used as experimental blanks. After incubation

154

for 0, 12, and 36 h, culture medium in each of wells was transferred to amber glass vials. For comparison on

155

EHDPP amounts between CEH cell and medium sample, the remaining cell layer in each of wells was also

156

collected by washing twice with 100 μL of ethanol, and transferred into amber glass vials. Prior to chemical

157

analysis, the collected medium or the ethanol/cell mixture was diluted by 25 times with fresh methanol. Then,

158

a 1000 µL aliquot of diluted samples were spiked with 20 ng of d15-TPHP (as internal standard for

159

quantification of EHDPP) and d10-DPHP (as internal standard for quantification of DPHP), filter-centrifuged,

160

and ready for instrumental analysis. ACS Paragon Plus Environment

Page 9 of 34

Environmental Science & Technology

161 162

Extraction of EHDPP Metabolites from Blood Samples

163

Screening of EHDPP metabolites was conducted in a single human blood pool comprising blood samples

164

collected during routine physical examinations of n = 99 people (50 males and 49 females) from Jiangsu

165

Province (cities of Suzhou, Taizhou, Yangzhou and Nanjing), eastern China. The age of recruited people

166

ranged from 18 to 87 years old. This research was approved by the research ethics boards of involved

167

universities, and the informed consent process has been conducted for all the volunteers prior to blood

168

sampling. An aliquot of 100 μL was taken from each of the 99 samples, and pooled together for further

169

extraction. Prior to extraction, 0.5 mL of blood sample was transferred into a disposable plain conical

170

centrifuge tube, and spiked with 10 ng of d15-TPHP and d10-DPHP. After that, 1.5 g Na2SO4, 0.1 g NaCl and

171

2 mL acetonitrile containing 1 % acetic acid were added to the sample. The sample tube was vortexed for 20

172

s and centrifuged for 5 min at 3500 rpm. A total volume of 1.5 mL of the supernatant fraction was transferred

173

into a disposable plain conical centrifuge tube, and 2.0 g Na2SO4 was added to the tube. The tube was vortexed

174

and centrifuged as described above. An aliquot of 1.0 mL of the supernatant fraction was transferred into a

175

new tube. The extract was evaporated under a stream of nitrogen to dryness, and 0.2 mL of methanol was

176

added to the sample. The sample was vortexed and was then ready for injection into the analytical instruments.

177

During the extraction process, recoveries of d10-DPHP and d15-TPHP were calculated to be 87 ± 8% and 93 ±

178

7% by comparing the instrumental differences of the spiked extracts and pure standards, respectively,

179

indicating a qualitative extraction of EHDPP metabolites from the plasma samples.

180 181

Liquid Chromatography-Quadrupole Mass Spectrometry (LC-MS/MS)

182

LC-ESI-MS/MS was used for: 1) the accurate quantification of EHDPP and DPHP in medium and cell

183

samples; and 2) investigation of daughter ions of newly detected EHDPP metabolites. For more details

184

regarding analysis of EHDPP or DPHP, please refer to our previous publications.5, 25 In brief, analysis of these ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 34

185

three compounds has been conducted by use of a high performance liquid chromatography (LC) coupled to a

186

tandem quadrupole mass spectrometer (MS/MS) (Waters, Milford, MA, U.S.A.). Chemicals were separated

187

by a C18 column (Waters Symmetry, 100 × 2.1 mm, particle size: 3.5 μm). The LC mobile phases were water

188

(containing 2 mM of ammonium acetate (AA)) and methanol (containing 2 mM of AA) with a constant flow

189

rate of 0.5 mL/min. The gradient of mobile phase was set as: 0 min, 5% methanol (containing 2 mM of AA);

190

0−5 min, 95% methanol (containing 2 mM of AA); hold for 1 min; 6−6.1 min, 5% methanol (containing 2

191

mM of AA) and hold for 4.9 min. For detection of EHDPP, the electrospray ionization (ESI) source was

192

operated in multiple reaction monitoring (MRM) mode and in positive mode, whereas for DPHP the ESI

193

source was operated in MRM mode and in negative mode. For OH-EHMPP metabolite, the ESI source was

194

operated in daughter ion mode and in negative mode to investigate formed daughter ions from potential OH-

195

EHMPP metabolites.

196 197

Liquid Chromatography-Quadrupole-Time of Flight-Mass Spectrometry (LC-Q-ToF-MS)

198

LC-Q-ToF/MS was used for screening potential EHDPP metabolites in CEHs and human blood samples.

199

The LC-Q-ToF/MS is an Agilent 1200 LC system that was coupled with an 6520A Q-ToF-MS system (Agilent

200

Technologies, Mississauga, ON, Canada). The Phenomenex (Torrance, CA, U.S.A.) Luna C18 column (2.0 ×

201

50 mm, particle size: 3 μm) has been used for separation of target compounds. The LC mobile phases were

202

water (containing 2 mM of ammonium acetate (AA)) and methanol (containing 2 mM of AA) with a constant

203

flow rate of 0.5 mL/min. The gradient for mobile phases was set as: 0 min, 5 % methanol (containing 2 mM

204

of AA); 0−5 min, 95 % methanol (containing 2 mM of AA); and hold for 10 min; and finally with a post run

205

time of 15 min. The LC-Q-ToF-MS was equipped with an ESI source, and each of the samples were injected

206

twice with ESI operated in either positive or negative mode. Before instrumental analysis, the high resolution

207

Q-ToF system was tuned and calibrated. Specifically, the resolution of Q-ToF system was >20,000 at m/z

208

322.0481 (Δ = 0.01 ppm) or m/z 301.9981 (Δ = 0.15 ppm) when the ESI interface was operated in positive or ACS Paragon Plus Environment

Page 11 of 34

Environmental Science & Technology

209

negative mode, respectively. For negative ESI, the reagent containing TFA anion (theoretical m/z 112.9855)

210

and HP-0921 formate adduct (theoretical m/z 966.0007) were consistently introduced into the Q-ToF-MS as

211

reference masses. For positive ESI, purine (theoretical m/z 121.0509) and HP-0921 (theoretical m/z 922.0098)

212

were used as reference masses.

213 214

Data Analysis

215

For cell viability, luciferase activity for each sample was adjusted for background (300 μM TDCIPP)

216

prior to the normalization as compared to the solvent control (DMSO). By use of GraphPad 5 (San Diego,

217

CA, U.S.A.), a nonlinear regression curve has been fitted for visualization of cell viability of EHDPP. PCR

218

array data were analyzed by use of the MxPro software (v4.10, Agilent Technologies, Santa Clara, CA,

219

U.S.A.), and the Ct was set to 0.1. A classic 2−ΔΔCt method has been used for calculation of fold change mRNA

220

expression compared to the vehicle control.36 Then, an one-way ANOVA (p < 0.05) analysis was used for

221

investigation of differences in fold change as compared to DMSO. For visualization of gene expression, target

222

genes with nonsignificant fold-changes (i.e. p > 0.05) or those less than 2-fold were set to 0 to minimize noise.

223

Finally, gene clustering and heatmaps were generated by use of a “gplots” package in R software (version:

224

3.0.2). For concentrations of EHDPP and DPHP in CEH, the peak in the chromatogram with an S/N ratio of

225

10 and 3 was identified as the method limit of quantification (MLOQ) and method limit of detection (MLOD),

226

respectively. MLOQ and MLOD for EHDPP were 0.03 and 0.01 ng/mL, respectively. MLOQs and MLODs

227

for DPHP were 0.6 and 0.2 ng/mL, respectively. Concentrations lower than MLODs, and those between

228

MLOQs and MLODs were reported as not detected (ND) and < MLOQ, respectively.

229 230

Results and discussion

231

Cytotoxic Effect Following Exposure to EHDPP

232

Cell viability was evaluated by measuring the cytoplasmic ATP levels in CEH exposed to a wide range ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 34

233

of EHDPP concentrations (0.1 μM to 100 μM). Any form of cell damage/death caused by EHDPP would be

234

reflected by lower concentrations of cytoplasmic ATP.37 A statistically significant decrease in the viability

235

(ANOVA, p < 0.05) was observed at exposure concentrations ≥ 10 μM EHDPP. Based on the fitted curve

236

(Figure 1A), the lethal concentration 50 (LC50) was calculated to be 50 ± 11 μM. To our knowledge, this is

237

the first report regarding cytotoxic effects of EHDPP in CEH. In a recent study, Hu et al investigated the

238

cytotoxicity of EHDPP following exposure to the human placental choriocarcinoma cell line, JEG-3. The

239

authors did not observe any effects on cell viability or proliferation at concentrations ranging from 5 to 40

240

μM.20 The observed differences might be due to the different cell lines used or species-specific responses.

241

Cell death was identified as a useful endpoint to prioritize chemical contaminants in vitro, and to date,

242

the CEH assay has been used to evaluate the cytotoxicity of 19 OPEs (Table S2).30, 33-35, 38 Including the present

243

study with EHDPP, of the 20 OPEs evaluated, 13 did not decrease cell viability up to the highest exposure

244

concentration. Alternatively, significant decreases in viability were observed for the following 6 OPEs as well

245

as EHDPP (50 ± 11 μM): tris(1,3-dichloro-2-propyl) phosphate (TDCIPP, LC50: 60.3 ± 45.8 μM), tris(2-

246

butoxyethyl) phosphate (TBOEP, 61.7 ± 43 μM), triphenyl phosphate (TPHP, 47 ± 8 μM), bis(2-ethylhexyl)

247

phosphate (DEHP, 40.6 ± 13 μM), p-tert-butylphenyl diphenyl phosphate (BPDP, 148 ± 83 μM), and

248

isopropylphenyl phosphate (IPPP, 179 ± 100 μM).26, 28-30, 33 Taken together, the LC50 of EHDPP is comparable

249

to a small cohort of highly cytotoxic OPEs, i.e. DEHP, TPHP, TDCIPP and TBOEP (Figure 1B), indicating

250

that EHDPP should be prioritized for further evaluation.

251 252

Alteration of mRNA Expression in CEH Exposed to EHDPP

253

The mRNA expression levels of 43 genes associated with the 9 pathways represented on the ToxChip

254

were examined in CEH following exposure to EHDPP at concentrations of 0.1, 1 or 10 μM. At least one gene

255

was altered in each of the pathways (Figure 2, Table S3) and 6, 11, or 16/43 genes were significantly

256

dysregulated following exposure to 0.1, 1, and 10 μM EHDPP, respectively. The greater transcriptomic ACS Paragon Plus Environment

Page 13 of 34

Environmental Science & Technology

257

response observed in CEHs treated with 10 μM demonstrated a concentration-dependent effect on gene

258

expression; individual fold-change values for those genes dysregulated by all treatment groups also followed

259

that pattern (Figure 2, Table S3).

260

All three genes from the ToxChip associated with bile acids/cholesterol regulation (CYP7B1, FGF19 and

261

NR5A2) were dysregulated by EHDPP; CYP7B1 and FGF19 were down-regulated and NR5A2 was up-

262

regulated. These genes play an important role in digestion and absorption of lipids in the small intestine.39 As

263

a cytochrome P450 enzyme associated with cholesterol synthesis, CYP7B1 catalyzes the conversion of 25-

264

hydroxycholesterol and 27-hydroxycholesterol to bile acid intermediates,40 and its inactivation could cause a

265

congenital bile acid synthesis deficiency in organisms.41 FGF19 , in combination with FGF receptor 4

266

(FGFR4), involved with the regulation of bile acid biosynthesis, gallbladder filling, and glucose and lipid

267

metabolism.42 NR5A2, also known as the liver receptor homolog-1 (LRH-1), plays a predominant role in

268

development, reverse cholesterol transport and bile acid homeostasis.43 Its expression could activate the

269

transcription of CYP7A1, which catalyzes the rate-limiting step in bile acid biosynthesis.44 Overall, genes

270

associated with bile acids/cholesterol homeostasis were altered by EHDPP, highlighting a novel mechanism

271

of action that should be further explored.

272

Both genes associated with the glucose metabolism pathway (PDK4 and G6pc) were down-regulated

273

following exposure to EHDPP. Glucose metabolism is critical for energy generation within the cell and the

274

supply of fuel for respiration.45 PDK4 is associated with the conversion of glucose to fatty acid and its

275

expression could be modulated by glucocorticoids (up-regulated) and insulin (down-regulated).46 PDK4 could

276

decrease the conversion of glucose to acetyl-CoA, which is associated with energy generation and maintaining

277

the balance between carbohydrate and lipid metabolism.47 G6pc is involved with maintenance of glucose

278

homeostasis by gluconeogenesis and glycogenolysis and its transcription could lead to lower glucose levels

279

by gluconeogenesis, which could be regulated by genes associated with energy homeostasis.48, 49 Collectively,

280

down-regulation of these two genes could result in a glucose metabolism imbalance and this adverse effect ACS Paragon Plus Environment

Environmental Science & Technology

281

Page 14 of 34

associated with EHDPP exposure requires further investigation.

282

Two of the 4 genes associated with lipid homeostasis (ACSL5 and SLCO1A2) were significantly down-

283

regulated in CEH following exposure to EHDPP. ACSL5 is a synthetase involved with the activation of fatty

284

acids to acyl-CoAs and plays an anabolic role in lipid metabolism with high rates of triacylglycerol synthesis.50,

285

51

286

uptake of a wide range of drugs and xenobiotics.52 Additional substrates (e.g. bile acids, steroids, thyroid

287

hormones and their conjugates) can also be transported by SLCO1A2.53 The concomitant down-regulation of

288

ACSL5 and SLCO1A2 could imply a disruption of lipid metabolism and energy storage by EHDPP. Alteration

289

of sphingolipid homeostasis by OPE-FRs was also reported in previous studies.54-56 Specifically, Du et al and

290

Morris et al reported that TPHP could disturb lipid metabolism in vitro (i.e. zebrafish liver) or in vivo (i.e.

291

mice), demonstrating a concordant endpoint in different model systems.55, 56 In a recent study, Zhao et al

292

reported that ln-transformed sphingolipid levels were negatively correlated with EHDPP concentrations in

293

human blood.54

SLCO1A2 encodes an organic anion-transporting polypeptide, which is an important mediator of cellular

294

Two of the 4 genes on the ToxChip associated with the thyroid hormone pathway (TTR and THRSP)

295

were down-regulated in a concentration-dependent manner following exposure to increasing concentrations

296

of EHDPP. The thyroid hormone pathway plays an important role in normal central nervous system

297

development in avian species, among many other critical roles.57 TTR is a major thyroid hormone carrier – in

298

addition to albumin – in avian species that transports thyroid hormones to target tissues.58 A depletion of

299

available thyroid hormones via decreased transport could lead to reduced total serum T4 levels.59-62 In addition

300

to EHDPP, other BFRs and OPEs (e.g. PBDEs, TBBPA, TPHP) have been identified as potent competitors

301

for thyroxine binding to TTR.9,

302

lipogenesis and maintenance of triacylglycerol and medium-chain fatty acid levels.63,64 To the best of our

303

knowledge, in vivo studies evaluating the effects of EHDPP on the thyroid hormone pathway are currently

304

unavailable. Given the importance of the thyroid hormone pathway for avian species, its disturbance following

65

THRSP is a crucial protein involved with thyroid hormone-mediated

ACS Paragon Plus Environment

Page 15 of 34

Environmental Science & Technology

305

exposure to various FRs represents an important research gap to further explore; in particular, via in vivo

306

studies.

307

Finally, across the other biological pathways represented on the ToxChip, multiple genes were responsive

308

to EHDPP exposure and provided additional insight with regards to the mechanism(s) of toxicity. For example,

309

genes from the cell cycle (GADD45a and DDB2), immune response (LEAP2), oxidative stress (OGG1) and

310

xenobiotic metabolism (SULT1B1 and FGA) pathways were dysregulated. Detailed network/pathway analysis

311

might elucidate additional information regarding the observed perturbations in CEHs.

312 313

EHDPP Degradation versus DPHP Formation in CEHs

314

In vitro metabolism of EHDPP was investigated by determining concentrations of EHDPP and its

315

potential metabolite, DPHP, in both medium and CEHs. We observed a rapid degradation of EHDPP in

316

medium and cells (Table 1 and Figure 3). Specifically, total concentrations (including both medium and cell

317

samples) of EHDPP at 0, 12 and 36 h were 0.96 ± 0.11, 0.33 ± 0.023 and 0.072 ± 0.008 μM, respectively

318

(Table 1). Concomitantly, total concentrations of DPHP at the same time points were ND, 0.1 ± 0.005 and

319

0.11 ± 0.02 μM, respectively (Table 1). At 36 h, approximately 92.5 % of the original EHDPP was depleted;

320

however, DPHP concentrations were only approximately 12 % of the original EHDPP concentration (Figure

321

3). A previous CEH in vitro TPHP metabolism study reported similar patterns regarding TPHP degradation

322

and DPHP formation.30 Specifically, by use of the same in vitro CEH assay, Su et al observed rapid

323

degradation of TPHP (i.e. 98 % of TPHP was depleted by 36 h) and the resulting concentrations of DPHP

324

accounted for 17 % of the initial TPHP dosing concentration.30 Taken together, these results suggest: 1) both

325

EHDPP and TPHP are susceptible to rapid degradation in CEHs; and 2) DPHP represents only one of the

326

metabolites of EHDPP or TPHP.

327

Based on the connection between cellular accumulation and specific potency, cell enrichment of EHDPP

328

and DPHP were determined in this study. Concentrations of EHDPP in CEH were 0.57 ± 0.03, 0.19 ± 0.02 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 34

329

and 0.042 ± 0.007 μM at time points of 0, 12 and 36 h, which represented 59 %, 57 % and 58 %, respectively,

330

of the total EHDPP concentration in wells. DPHP, on the other hand, was only detected in medium and not

331

CEHs at the two latter time points (12 and 36h). The Log octanol-water partition coefficient (Log KOW) of

332

EHDPP is 5.73 compared to 2.88 for DPHP, and therefore could help explain the preferential bioaccumulation

333

and prominent effects (e.g. cell viability and gene expression) observed for EHDPP.2, 66 Overall, the variable

334

distribution of EHDPP and DPHP in medium and cells indicated that: a) EHDPP was taken up by CEH; b)

335

DPHP was formed as a metabolite of EHDPP; and c) the metabolite, DPHP, was eliminated completely from

336

cells into the medium. These assumptions are consistent with the detection of DPHP in human fluids (i.e.,

337

urine) in previous studies.67, 68

338

DPHP has typically been regarded as a biomarker for measuring TPHP exposure in humans with frequent

339

detection in human urine samples from various countries around the world.25,

340

publications are raising questions regarding this link because: 1) DPHP can be formed from multiple parent

341

compounds (e.g. TPHP, EHDPP, RDP, ITP isomers, or TBPP isomers); and 2) a limited amount

342

(approximately 20 % or less) of DPHP was formed from TPHP via various in vitro assays that used human

343

microsomes, herring gull microsomes, CEHs or human serum.28, 30, 31, 70, 71 The results of the present study

344

provide additional evidence that DPHP is not a specific biomarker of TPHP given its detection in culture

345

medium following EHDPP exposure. Therefore, caution regarding the interpretation of DPHP detection is

346

recommended.

68, 69

However, recent

347 348

Comparison of EHDPP Metabolism in CEHs and humans

349

Previous in vitro studies, using human liver microsomes or S9 fractions, demonstrated that

350

biotransformation of OPEs can generate a wide array of metabolites formed by oxidative metabolism.70

351

Typically, microsomal enzymes (phase I) cause hydroxylation, O-dealkylation, oxidative diarylation and

352

reductive and oxidative dehalogenation, whereas S9 fractions (phase II) typically induce glucuronidation and ACS Paragon Plus Environment

Page 17 of 34

Environmental Science & Technology

353

sulfation of metabolites generated during phase I metabolism.70, 72-74 On this basic knowledge, we created a

354

database that includes 19 potential EHDPP metabolites, all of which were further screened in the EHDPP-

355

exposed CEH samples by high resolution mass spectrometry (Table S4).

356

In addition to DPHP, we detected an ion of m/z 301.1214 in CEHs, which was absent in the blanks (Figure 4).

357

The m/z 301.1214 was presumed to be hydroxylated 2-ethylhexyl monophenyl phosphate (OH-EHMPP) with

358

a theoretical formula of C14H23O5P1 with [M-H]- of m/z 301.1210. Using the MS/MS mode of a UPLC-ESI-

359

TQ-S/MS operated in negative mode, we further examined the fragmentation of m/z 301, which can generate

360

multiple daughter ions including m/z 79, m/z 93 and m/z 173. We presumed that the m/z 79 daughter ion was

361

generated from a resonance stabilization of m/z 301 via a McLafferty rearrangement. The m/z 79 ion is a

362

phosphite anion ([PO3]-), which has been reported for other diesters, i.e. DEHP and DNBP, in previous

363

studies.25, 75 The daughter ions m/z 93 and m/z 173 were presumed to be phenolate anion ([C6H5O1]-) and

364

mono-phenyl phosphate anion ([C6H6P1O4]-), respectively, which are typical daughter ions for aryl phosphate

365

esters, i.e. DPHP.75 Collectively, our datasets based on both high resolution Q-ToF/MS and MS/MS suggested

366

that: 1) the newly detected ion could be OH-EHMPP; and 2) hydroxylation likely occurred on the functional

367

group of 2-ethylhexyl. However, due to limited authentic and pure analytical standards, accurate and thorough

368

identification and quantitation of this new metabolite was not possible at present. To the best of our knowledge,

369

this is the first report describing the detection of OH-EHMPP in a CEH metabolism study. In previous studies,

370

the alkyl chain of EHDPP tended to undergo oxidation as opposed to the phenyl rings in human liver

371

microsomes.24 In contrast, the presence of OH-EHMPP revealed a hydroxylation metabolism pathway for

372

EHDPP, and similar hydroxylation metabolism has been shown for flame retardants including TPHP and

373

HBCDD.66, 76, 77

374

To further elucidate the differences of EHDPP metabolism in CEH versus human samples, the 19

375

potential EHDPP metabolites were also screened in a human blood pool collected from individuals from

376

Jiangsu Province, eastern China. Three metabolites, including DPHP, OH-EHMPP and EHMPP, were ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 34

377

detected in the human blood pool and EHMPP showed the greatest instrumental intensity (Figure 5). These

378

results suggested that: 1) the formation of OH-EHMPP from EHDPP could occur in both humans and avian

379

species; and 2) differences exist between human and avian species because EHMPP was not detectable in

380

CEH medium but showed greatest intensity in human blood samples. The detection of EHMPP and DPHP in

381

human samples is well supported by several recent studies.32,

382

pooled human serum, van den Eede et al detected the formation of EHMPP with a minor amount of DPHP.71

383

By screening 14 OPE metabolites in urine samples of 128 elementary school-aged children, Araki et al

384

reported that EHMPP was detectable in 65 % of analyzed samples, and DPHP was detectable in 80 % of

385

analyzed samples with a mean concentration of 0.51 ng/mL.32 Currently, there is no available information

386

regarding DPHP or EHMPP concentrations in human blood.

387 388 389

Acknowledgements:

71

Specifically, by incubating EHDPP with

390

This research was supported by “Natural Science Foundation of Jiangsu Province” (Grant No.

391

BK20170830 and BK20180498), “the Fundamental Research Funds for the Central Universities” (Grant No.

392

30917011305), and “Jiangsu Key Laboratory of Chemical Pollution Control and Resources Reuse (Nanjing

393

University of Science and Technology)” (Grant No. 30918014102). Dr. Su appreciates support from the

394

programs of “Thousand Talents Plan” and “Jiangsu Provincial Distinguished Professorship”.

395 396 397

Supporting Information The supporting information is available free of charge on the ACS Publications website. Further details

398

are given on biological function description of 43 target genes, a summary of LC50 values of 19 OPEs,

399

specific fold changes and p-values of 43 target genes following exposure to EHDPP, and a list of 19

400

potential EHDPP metabolites.

401 402 ACS Paragon Plus Environment

Page 19 of 34

Environmental Science & Technology

403

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 34

404

References:

405

1.

406

http://chm.pops.int/Convention/ThePOPs/TheNewPOPs/tabid/2511/Default.aspx. Accessed on September,

407

2018.

408

2.

409

occurrence, toxicity and analysis. Chemosphere 2012, 88, (10), 1119-1153.

410

3.

411

Q., Organophosphorus flame retardants and plasticizers: sources, occurrence, toxicity and human exposure.

412

Environ Pollut 2015, 196, 29-46.

413

4.

414

in Caspian tern compared to herring gull eggs from Michigan colonies in the Great Lakes of North America.

415

Environ Pollut 2017, 222, 154-164.

416

5.

417

dependent hydrolysis, kinetics, and pathways. Environ Sci Technol 2016, 50, (15), 8103-8111.

418

6.

419

Benedek Plosz, Mikael Remberger,; Merete Schøyen, H. T. U., Christian Vogelsang, Screening of selected

420

metals and new organic contaminants 2007.

421

7.

422

diphenyl phosphate (CAS no. 1241-94-7). 2009.

423

8.

424

halogenated flame retardants, and polybrominated diphenyl ethers in indoor and outdoor air in Stockholm,

425

Sweden. Environ Pollut 2018, 240, 514-522.

426

9.

427

and their transfer to chorionic villi. Environ Sci Technol 2017, 51, (11), 6489-6497.

Stockholm

Convention;

The

new

POPs

under

the

Stockholm

Convention;

Website:

van der Veen, I.; de Boer, J., Phosphorus flame retardants: Properties, production, environmental

Wei, G. L.; Li, D. Q.; Zhuo, M. N.; Liao, Y. S.; Xie, Z. Y.; Guo, T. L.; Li, J. J.; Zhang, S. Y.; Liang, Z.

Su, G.; Letcher, R. J.; Moore, J. N.; Williams, L. L.; Grasman, K. A., Contaminants of emerging concern

Su, G.; Letcher, R. J.; Yu, H., Organophosphate flame retardants and plasticizers in aqueous solution: pH-

Norman Green (project leader), M. S.; Torgeir Bakke, E. M. B., Christian Dye, Dorte; Herzke, S. H.,

Brooke D N, C. M. J., Quarterman P and Burns J, Environmental risk evaluation report: 2-Ethylhexyl

Wong, F.; de Wit, C. A.; Newton, S. R., Concentrations and variability of organophosphate esters,

Zhao, F.; Chen, M.; Gao, F.; Shen, H.; Hu, J., Organophosphorus flame retardants in pregnant women

ACS Paragon Plus Environment

Page 21 of 34

Environmental Science & Technology

428

10. Kim, U. J.; Oh, J. K.; Kannan, K., Occurrence, removal, and environmental emission of organophosphate

429

flame retardants/plasticizers in a wastewater treatment plant in New York State. Environ Sci Technol 2017,

430

51, (14), 7872-7880.

431

11. Phillips, A. L.; Hammel, S. C.; Hoffman, K.; Lorenzo, A. M.; Chen, A.; Webster, T. F.; Stapleton, H. M.,

432

Children's residential exposure to organophosphate ester flame retardants and plasticizers: Investigating

433

exposure pathways in the TESIE study. Environ Int 2018, 116, 176-185.

434

12. Greaves, A. K.; Letcher, R. J., Comparative body compartment composition and in ovo transfer of

435

organophosphate flame retardants in north american great lakes herring gulls. Environ Sci Technol 2014, 48,

436

(14), 7942-7950.

437

13. Coelho, S. D.; Sousa, A. C. A.; Isobe, T.; Kim, J. W.; Kunisue, T.; Nogueira, A. J. A.; Tanabe, S.,

438

Brominated, chlorinated and phosphate organic contaminants in house dust from Portugal. Sci Total Environ

439

2016, 569-570, 442-449.

440

14. Langer, S.; Fredricsson, M.; Weschler, C. J.; Beko, G.; Strandberg, B.; Remberger, M.; Toftum, J.;

441

Clausen, G., Organophosphate esters in dust samples collected from Danish homes and daycare centers.

442

Chemosphere 2016, 154, 559-566.

443

15. Brommer, S.; Harrad, S., Sources and human exposure implications of concentrations of organophosphate

444

flame retardants in dust from UK cars, classrooms, living rooms, and offices. Environ Int 2015, 83, 202-7.

445

16. Liu, Y.; Song, N.; Guo, R.; Xu, H.; Zhang, Q.; Han, Z.; Feng, M.; Li, D.; Zhang, S.; Chen, J., Occurrence

446

and partitioning behavior of organophosphate esters in surface water and sediment of a shallow Chinese

447

freshwater lake (Taihu Lake): Implication for eco-toxicity risk. Chemosphere 2018, 202, 255-263.

448

17. Xing, L.; Zhang, Q.; Sun, X.; Zhu, H.; Zhang, S.; Xu, H., Occurrence, distribution and risk assessment of

449

organophosphate esters in surface water and sediment from a shallow freshwater Lake, China. Sci Total

450

Environ 2018, 636, 632-640.

451

18. Kim, J. W.; Isobe, T.; Chang, K. H.; Amano, A.; Maneja, R. H.; Zamora, P. B.; Siringan, F. P.; Tanabe, ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 34

452

S., Levels and distribution of organophosphorus flame retardants and plasticizers in fishes from Manila Bay,

453

the Philippines. Environ Pollut 2011, 159, (12), 3653-3659.

454

19. Sundkvist, A. M.; Olofsson, U.; Haglund, P., Organophosphorus flame retardants and plasticizers in

455

marine and fresh water biota and in human milk. J Environ Monitor 2010, 12, (4), 943-951.

456

20. Hu, W.; Gao, F.; Zhang, H.; Hiromori, Y.; Arakawa, S.; Nagase, H.; Nakanishi, T.; Hu, J., Activation of

457

peroxisome proliferator-activated receptor gamma and disruption of progesterone synthesis of 2-ethylhexyl

458

diphenyl phosphate in human placental choriocarcinoma cells: comparison with triphenyl phosphate. Environ

459

Sci Technol 2017, 51, (7), 4061-4068.

460

21. Cao, L.-Y.; Ren, X.-M.; Li, C.-H.; Guo, L.-H., Organophosphate esters bind to and inhibit estrogen-

461

related receptor γ in cells. Environ Sci Technol Lett 2018, 5, (2), 68-73.

462

22. Schang, G.; Robaire, B.; Hales, B. F., Organophosphate flame retardants act as endocrine-disrupting

463

chemicals in ma-10 mouse tumor leydig cells. Toxicol Sci 2016, 150, (2), 499-509.

464

23. Behl, M.; Hsieh, J. H.; Shafer, T. J.; Mundy, W. R.; Rice, J. R.; Boyd, W. A.; Freedman, J. H.; Hunter,

465

E. S., 3rd; Jarema, K. A.; Padilla, S.; Tice, R. R., Use of alternative assays to identify and prioritize

466

organophosphorus flame retardants for potential developmental and neurotoxicity. Neurotoxicol Teratol 2015,

467

52, 181-93.

468

24. Ballesteros-Gomez, A.; Erratico, C. A.; Eede, N. V.; Ionas, A. C.; Leonards, P. E.; Covaci, A., In vitro

469

metabolism of 2-ethylhexyldiphenyl phosphate (EHDPHP) by human liver microsomes. Toxicol Lett 2015,

470

232, (1), 203-12.

471

25. Su, G.; Letcher, J. R.; Yu, H., Determination of organophosphate diesters in urine samples by a high-

472

sensitivity method based on ultra high pressure liquid chromatography-triple quadrupole-mass spectrometry.

473

J Chromatogr A 2015, 1426, 154-160.

474

26. Hoffman, K.; Daniels, J. L.; Stapleton, H. M., Urinary metabolites of organophosphate flame retardants

475

and their variability in pregnant women. Environ Int 2014, 63, 169-72. ACS Paragon Plus Environment

Page 23 of 34

Environmental Science & Technology

476

27. Preston, E. V.; McClean, M. D.; Claus Henn, B.; Stapleton, H. M.; Braverman, L. E.; Pearce, E. N.;

477

Makey, C. M.; Webster, T. F., Associations between urinary diphenyl phosphate and thyroid function. Environ

478

Int 2017, 101, 158-164.

479

28. Ballesteros-Gomez, A.; Van den Eede, N.; Covaci, A., In vitro human metabolism of the flame retardant

480

resorcinol bis(diphenylphosphate) (RDP). Environ Sci Technol 2015, 49, (6), 3897-904.

481

29. Su, G.; Letcher, R. J.; Crump, D.; Gooden, D. M.; Stapleton, H. M., In vitro metabolism of the flame

482

retardant triphenyl phosphate in chicken embryonic hepatocytes and the importance of the hydroxylation

483

pathway. Environ Sci Technol Lett 2015, 2, (4), 100-104.

484

30. Su, G.; Crump, D.; Letcher, R. J.; Kennedy, S. W., Rapid in vitro metabolism of the flame retardant

485

triphenyl phosphate and effects on cytotoxicity and mRNA expression in chicken embryonic hepatocytes.

486

Environ Sci Technol 2014, 48, (22), 13511-9.

487

31. Phillips, A. L.; Hammel, S. C.; Konstantinov, A.; Stapleton, H. M., Characterization of individual

488

isopropylated and tert-butylated triarylphosphate (ITP and TBPP) isomers in several commercial flame

489

retardant mixtures and house dust Standard Reference Material SRM 2585. Environ Sci Technol 2017, 51,

490

(22), 13443-13449.

491

32. Araki, A.; Bastiaensen, M.; Ait Bamai, Y.; Van den Eede, N.; Kawai, T.; Tsuboi, T.; Ketema, R. M.;

492

Covaci, A.; Kishi, R., Associations between allergic symptoms and phosphate flame retardants in dust and

493

their urinary metabolites among school children. Environ Int 2018, 119, 438-446.

494

33. Porter, E.; Crump, D.; Egloff, C.; Chiu, S.; Kennedy, S. W., Use of an avian hepatocyte assay and the

495

avian Toxchip Polymerse chain reaction array for testing prioritization of 16 organic flame retardants. Environ

496

Toxicol Chem 2014, 33, (3), 573-82.

497

34. Page-Lariviere, F.; Chiu, S.; Jones, S. P.; Farhat, A.; Crump, D.; O'Brien, J. M., Prioritization of 10

498

organic flame retardants using an avian hepatocyte toxicogenomic assay. Environ Toxicol Chem 2018, 37,

499

(12): 3134-3144. ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 34

500

35. Crump, D.; Chiu, S.; Kennedy, S. W., Effects of tris(1,3-dichloro-2-propyl) phosphate and tris(1-

501

chloropropyl) phosphate on cytotoxicity and mRNA expression in primary cultures of avian hepatocytes and

502

neuronal cells. Toxicol Sci 2012, 126, (1), 140-8.

503

36. Zarandi, E. R.; Mansouri, S.; Nakhaee, N.; Sarafzadeh, F.; Moradi, M., Effect of sub-MIC of vancomycin

504

and clindamycin alone and in combination with ceftazidime on Clostridium difficile surface layer protein A

505

(slpA) gene expression. Microb Pathogenesis 2017, 111, 163-167.

506

37. Crouch, S. P. M.; Kozlowski, R.; Slater, K. J.; Fletcher, J., The use of ATP bioluminescence as a measure

507

of cell proliferation and cytotoxicity. J Immunol Methods 1993, (160), 81-88.

508

38. Egloff, C.; Crump, D.; Porter, E.; Williams, K. L.; Letcher, R. J.; Gauthier, L. T.; Kennedy, S. W., Tris(2-

509

butoxyethyl)phosphate and triethyl phosphate alter embryonic development, hepatic mRNA expression,

510

thyroid hormone levels, and circulating bile acid concentrations in chicken embryos. Toxicol Appl Pharmacol

511

2014, 279, (3), 303-310.

512

39. Makishima, M., Nuclear receptors as targets for drug development: regulation of cholesterol and bile acid

513

metabolism by nuclear receptors. J Pharmacol Sci 2005, 97, (2), 177-183.

514

40. Russell, D. W., The enzymes, regulation, and genetics of bile acid synthesis. Annu Rev Biochem 2003,

515

72, (1), 137-174.

516

41. Lorbek, G.; Lewinska, M.; Rozman, D., Cytochrome P450s in the synthesis of cholesterol and bile acids–

517

from mouse models to human diseases. FEBS J 2012, 279, (9), 1516-1533.

518

42. Jones, S., Mini-review: Endocrine actions of fibroblast growth factor 19. Mol Pharm 2008, 5 (1), 42-48.

519

43. Fayard, E.; Auwerx, J.; Schoonjans, K., LRH-1: an orphan nuclear receptor involved in development,

520

metabolism and steroidogenesis. Trends Cell Biol 2004, 14, (5), 250-260.

521

44. Goodwin, B.; Jones, S. A.; Price, R. R.; Watson, M. A.; McKee, D. D.; Moore, L. B.; Galardi, C.; Wilson,

522

J. G.; Lewis, M. C.; Roth, M. E., A regulatory cascade of the nuclear receptors FXR, SHP-1, and LRH-1

523

represses bile acid biosynthesis. Mol Cell 2000, 6, (3), 517-526. ACS Paragon Plus Environment

Page 25 of 34

Environmental Science & Technology

524

45. Randle, P.; Garland, P.; Hales, C.; Newsholme, E., The glucose fatty-acid cycle its role in insulin

525

sensitivity and the metabolic disturbances of diabetes mellitus. Lancet 1963, 281, (7285), 785-789.

526

46. Connaughton, S.; Chowdhury, F.; Attia, R. R.; Song, S.; Zhang, Y.; Elam, M. B.; Cook, G. A.; Park, E.

527

A., Regulation of pyruvate dehydrogenase kinase isoform 4 (PDK4) gene expression by glucocorticoids and

528

insulin. Mol Cell Endocrinol 2010, 315, (1-2), 159-167.

529

47. Wang, D.; Zhu, W.; Chen, L.; Yan, J.; Teng, M.; Zhou, Z., Neonatal triphenyl phosphate and its metabolite

530

diphenyl phosphate exposure induce sex- and dose-dependent metabolic disruptions in adult mice. Environ

531

Pollut 2018, 237, 10-17.

532

48. Chou, J. Y.; Mansfield, B. C., Mutations in the glucose‐6‐phosphatase‐α (G6PC) gene that cause type Ia

533

glycogen storage disease. Hum Mutat 2008, 29, (7), 921-930.

534

49. Guo, F.; Zhang, Y.; Zhang, C.; Wang, S.; Ni, Y.; Zhao, R., Fat mass and obesity associated (FTO) gene

535

regulates gluconeogenesis in chicken embryo fibroblast cells. Comp Biochem Physiol A Mol Integr Physiol

536

2015, 179, 149-156.

537

50. Bu, S. Y.; Mashek, D. G., Hepatic long-chain acyl-CoA synthetase 5 (ACSL5) mediates fatty acid

538

channeling between anabolic and catabolic pathways. J Lipid Res 2010, 51, (11), 3270-3280.

539

51. Bowman, T. A.; O'Keeffe, K. R.; D'aquila, T.; Yan, Q. W.; Griffin, J. D.; Killion, E. A.; Salter, D. M.;

540

Mashek, D. G.; Buhman, K. K.; Greenberg, A. S., Acyl CoA synthetase 5 (ACSL5) ablation in mice increases

541

energy expenditure and insulin sensitivity and delays fat absorption. Mol Metab 2016, 5, (3), 210-220.

542

52. Lee, W.; Glaeser, H.; Smith, L. H.; Roberts, R. L.; Moeckel, G. W.; Gervasini, G.; Leake, B. F.; Kim, R.

543

B., Polymorphisms in Human Organic Anion-transporting Polypeptide 1A2 (OATP1A2) implications for

544

altered drug disposition and central nervous system drug entry. J Biol Chem 2005, 280, (10), 9610-9617.

545

53. Zhou, F.; Zheng, J.; Zhu, L.; Jodal, A.; Cui, P. H.; Wong, M.; Gurney, H.; Church, W. B.; Murray, M.,

546

Functional analysis of novel polymorphisms in the human SLCO1A2 gene that encodes the transporter

547

OATP1A2. AAPS J 2013, 15, (4), 1099-1108. ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 34

548

54. Zhao, F. R.; Wan, Y.; Zhao, H. Q.; Hu, W. X.; Mu, D.; Webster, T. F.; Hu, J. Y., Levels of blood

549

organophosphorus flame retardants and association with changes in human sphingolipid homeostasis. Environ

550

Sci Technol 2016, 50, (16), 8896-8903.

551

55. Du, Z.; Zhang, Y.; Wang, G.; Peng, J.; Wang, Z.; Gao, S., TPhP exposure disturbs carbohydrate

552

metabolism, lipid metabolism, and the DNA damage repair system in zebrafish liver. Sci Rep 2016, 6, 21827.

553

56. Morris, P. J.; Medina-Cleghorn, D.; Heslin, A.; King, S. M.; Orr, J.; Mulvihill, M. M.; Krauss, R. M.;

554

Nomura, D. K., Organophosphorus flame retardants inhibit specific liver carboxylesterases and cause serum

555

hypertriglyceridemia. ACS Chem Biol 2014, 9, (5), 1097-103.

556

57. McNabb, F. A., The hypothalamic-pituitary-thyroid (HPT) axis in birds and its role in bird development

557

and reproduction. Crit Rev Toxicol 2007, 37, (1-2), 163-193.

558

58. Palha, J. A., Transthyretin as a thyroid hormone carrier: function revisited. Clin Chem Lab Med 2002, 40,

559

(12), 1292-1300.

560

59. Dussault, J. H.; Ruel, J., Thyroid hormones and brain development. Annu Rev Physiol 1987, 49, (1), 321-

561

332.

562

60. Episkopou, V.; Maeda, S.; Nishiguchi, S.; Shimada, K.; Gaitanaris, G. A.; Gottesman, M. E.; Robertson,

563

E. J., Disruption of the transthyretin gene results in mice with depressed levels of plasma retinol and thyroid

564

hormone. P Natl Acad Sci USA 1993, 90, (6), 2375-2379.

565

61. Porterfield, S. P.; Hendrich, C. E., The role of thyroid hormones in prenatal and neonatal neurological

566

development—current perspectives. Endocr Rev 1993, 14, (1), 94-106.

567

62. McNabb, F., Biomarkers for the assessment of avian thyroid disruption by chemical contaminants. Avian

568

Poult Biol Rev 2005, 16, (1), 3-10.

569

63. LaFave, L. T.; Augustin, L. B.; Mariash, C. N., S14: insights from knockout mice. Endocrinology 2006,

570

147, (9), 4044-4047.

571

64. Yao, D.; Luo, J.; He, Q.; Wu, M.; Shi, H.; Wang, H.; Wang, M.; Xu, H.; Loor, J., Thyroid hormone ACS Paragon Plus Environment

Page 27 of 34

Environmental Science & Technology

572

responsive (THRSP) promotes the synthesis of medium-chain fatty acids in goat mammary epithelial cells. J

573

Dairy Sci 2016, 99, (4), 3124-3133.

574

65. Meerts, I. A.; Van Zanden, J. J.; Luijks, E. A.; van Leeuwen-Bol, I.; Marsh, G.; Jakobsson, E.; Bergman,

575

Å.; Brouwer, A., Potent competitive interactions of some brominated flame retardants and related compounds

576

with human transthyretin in vitro. Toxicol Sci 2000, 56, (1), 95-104.

577

66. Greaves, A. K.; Su, G.; Letcher, R. J., Environmentally relevant organophosphate triesters in herring

578

gulls: In vitro biotransformation and kinetics and diester metabolite formation using a hepatic microsomal

579

assay, Toxicol Appl Pharmacol 2016, 308, 59-65.

580

67. Carignan, C. C.; Minguez-Alarcon, L.; Williams, P. L.; Meeker, J. D.; Stapleton, H. M.; Butt, C. M.;

581

Toth, T. L.; Ford, J. B.; Hauser, R.; Team, E. S., Paternal urinary concentrations of organophosphate flame

582

retardant metabolites, fertility measures, and pregnancy outcomes among couples undergoing in vitro

583

fertilization. Environ Int 2018, 111, 232-238.

584

68. Hoffman, K.; Butt, C. M.; Chen, A.; Limkakeng, A. T., Jr.; Stapleton, H. M., High exposure to

585

organophosphate flame retardants in infants: associations with baby products. Environ Sci Technol 2015, 49,

586

(24), 14554-9.

587

69. Mendelsohn, E.; Hagopian, A.; Hoffman, K.; Butt, C. M.; Lorenzo, A.; Congleton, J.; Webster, T. F.;

588

Stapleton, H. M., Nail polish as a source of exposure to triphenyl phosphate. Environ Int 2016, 86, 45-51.

589

70. Van den Eede, N.; Maho, W.; Erratico, C.; Neels, H.; Covaci, A., First insights in the metabolism of

590

phosphate flame retardants and plasticizers using human liver fractions. Toxicol Lett 2013, 223, (1), 9-15.

591

71. Van den Eede, N.; Ballesteros-Gomez, A.; Neels, H.; Covaci, A., Does Biotransformation of Aryl

592

Phosphate Flame Retardants in Blood Cast a New Perspective on Their Debated Biomarkers? Environ Sci

593

Technol 2016, 50, (22), 12439-12445.

594

72. Levi, P.; Hollingworth, R.; Hodgson, E., Differences in oxidative dearylation and desulfuration of

595

fenitrothion by cytochrome-P450 isozymes and in the subsequent inhibition of monooxygenase activity. ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 34

596

Pesticide Biochem Physiol 1988, 32, 224-231.

597

73. Testa, B.; Krämer, S. D., The biochemistry of drug metabolism: principles, redox Reactions, hydrolyses.

598

Verlag Helvetica Chimica Acta, Zürich, Switzerland 2010.

599

74. Testa, B.; Krämer, S. D., The biochemistry of drug metabolism: conjugations, consequences of

600

metabolism, influencing factors. Verlag Helvetica Chimica Acta, Zürich, Switzerland 2010.

601

75. Chu, S.; Chen, D.; Letcher, R. J., Dicationic ion-pairing of phosphoric acid diesters post-liquid

602

chromatography and subsequent determination by electrospray positive ionization-tandem mass spectrometry.

603

J Chromatogr A 2011, 1218, (44), 8083-8.

604

76. Antonopoulou, M.; Karagianni, P.; Konstantinou, I. K., Kinetic and mechanistic study of photocatalytic

605

degradation of flame retardant Tris (1-chloro-2-propyl) phosphate (TCPP). Appl Catal B-Environ 2016, 192,

606

152-160.

607

77. Esslinger, S.; Becker, R.; Maul, R.; Nehls, I., Hexabromocyclododecane enantiomers: microsomal

608

degradation and patterns of hydroxylated metabolites. Environ Sci Technol 2011, 45, (9), 3938-3944.

609 610

ACS Paragon Plus Environment

Page 29 of 34

611 612 613 614

615 616 617 618

Environmental Science & Technology

Table 1. Mean concentrations ± SD (units: μM) of 2-ethylhexyl diphenyl phosphate (EHDPP) and the metabolite diphenyl phosphate (DPHP) in cell culture medium and hepatocytes at three time points (0, 12 and 36 h) following exposure of chicken embryonic hepatocytes to EHDPP. Nominal concentration of EHDPP was 1 μM. Chemicals Samples 0h 12 h 36 h Medium 0.39 ± 0.08 0.14 ± 0.003 0.030 ± 0.001 EHDPP Cells 0.57 ± 0.03 0.19 ± 0.02 0.042 ± 0.007 % EHDPP in cells 59 % 57 % 58 % b Medium ND 0.10 ± 0.005 0.11 ± 0.02 DPHP Cells ND ND ND % DPHP in cells NCa NC NC a “NC” means not calculated because chemical was not detectable in cell or medium samples; b MLOQ and MLOD for EHDPP were 0.03 and 0.01 ng/mL, respectively. MLOQs and MLODs for DPHP were 0.6 and 0.2 ng/mL, respectively. Concentrations lower than MLODs and those between MLOQs and MLODs are reported as not detected (ND) and < MLOQ, respectively.

ACS Paragon Plus Environment

Environmental Science & Technology

619 620 621 622 623

Page 30 of 34

Figure 1. Concentration-dependent cytotoxicity of 2-ethylhexyl diphenyl phosphate (EHDPP) in chicken embryonic hepatocytes (CEHs) incubated for 36h (A) and lethal concentration 50 (LC50) ranking for various organophosphate esters (OPEs) based on the CEH assay (B). The data in Figure 1A were log-transformed and the percent cell viability data (n=3 well per treatment group) were fit to a nonlinear regression curve (log (agonist) vs response) to determine the LC50 (±SEM). Error bars for each point represent the standard deviation of three replicates. LC50 values for a total of 19 OPEs are summarized in Table S2, and only those that exhibited cytotoxicity up to the highest administered concentrations are shown in Figure 1B.

624 625

ACS Paragon Plus Environment

Page 31 of 34

626 627 628 629 630

Environmental Science & Technology

Figure 2. The relative mRNA expression profiles of 43 genes on the Avian ToxChip PCR array following exposure to 3 concentrations of 2-ethylhexyl diphenyl phosphate (EHDPP; 0.1 μM, 1 μM and 10 μM) for 36h. The heat map data represent the mean of three replicates. Genes with fold change lower than 2.0 and p > 0.05 were set to 0 to minimize noise. Green-black-red indicate down-, no-, and up-regulation. The detailed mRNA fold-change data are available in Table S3.

631 632

ACS Paragon Plus Environment

Environmental Science & Technology

633 634 635 636 637

Page 32 of 34

Figure 3. Mass balance between 2-ethylhexyl diphenyl phosphate (EHDPP; the parent compound) and the metabolite, diphenyl phosphate (DPHP), following exposure of chicken embryonic hepatocytes (CEHs) to 1 μM EHDPP for 0, 12, and 36h. Reported concentrations (µM) were total concentrations of chemicals in both medium and cell samples. The data represent the mean of three replicate wells per time point.

638 639

ACS Paragon Plus Environment

Page 33 of 34

640 641 642 643 644

Environmental Science & Technology

Figure 4. Detection of parent 2-ethylhexyl diphenyl phosphate (EHDPP) in culture medium at time point 0 h (A), diphenyl phosphate (DPHP) in culture medium at time point 36 h (B), and hydroxylated 2-ethylhexyl monophenyl phosphate (OH-EHMPP) in culture medium at time point 36 h (C) by use of liquid chromatography (LC)-electrospray (ESI)-time-of-flight mass spectrometry. Assignment of daughter ions (D) was generated from OH-EHMPP by use of LCESI-triple quadrupole mass spectrometry. For EHDPP, the ESI source was operated in positive mode, whereas the ESI source was operated in negative mode for DPHP and OH-EHMPP. “TM” means theoretical mass, and “ME” means error between theoretical and observed masses.

645 646

ACS Paragon Plus Environment

Environmental Science & Technology

647 648 649 650

Page 34 of 34

Figure 5. Metabolites of 2-ethylhexyl diphenyl phosphate (EHDPP) detected in a pooled human blood sample by use of liquid chromatography-electrospray (ESI)-time-of-flight mass spectrometry: A) diphenyl phosphate (DPHP), B) hydroxylated 2-ethylhexyl monophenyl phosphate (OH-EHMPP), and C) 2ethylhexyl monophenyl phosphate (EHMPP). For all three metabolites, the ESI source was operated in negative mode. “TM” means theoretical mass, and “ME” means error between theoretical and observed masses.

651 652 653

ACS Paragon Plus Environment